Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Environmental Inorganic Chemistry for Engineers
Environmental Inorganic Chemistry for Engineers
Environmental Inorganic Chemistry for Engineers
Ebook1,037 pages30 hours

Environmental Inorganic Chemistry for Engineers

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Environmental Inorganic Chemistry for Engineers explains the principles of inorganic contaminant behavior, also applying these principles to explore available remediation technologies, and providing the design, operation, and advantages or disadvantages of the various remediation technologies.

Written for environmental engineers and researchers, this reference provides the tools and methods that are imperative to protect and improve the environment. The book's three-part treatment starts with a clear and rigorous exposition of metals, including topics such as preparations, structures and bonding, reactions and properties, and complex formation and sequestering.

This coverage is followed by a self-contained section concerning complex formation, sequestering, and organometallics, including hydrides and carbonyls. Part Two, Non-Metals, provides an overview of chemical periodicity and the fundamentals of their structure and properties.

  • Clearly explains the principles of inorganic contaminant behavior in order to explore available remediation technologies
  • Provides the design, operation, and advantages or disadvantages of the various remediation technologies
  • Presents a clear exposition of metals, including topics such as preparations, structures, and bonding, reaction and properties, and complex formation and sequestering
LanguageEnglish
Release dateMay 10, 2017
ISBN9780128011423
Environmental Inorganic Chemistry for Engineers
Author

James G. Speight

Dr. Speight is currently editor of the journal Petroleum Science and Technology (formerly Fuel Science and Technology International) and editor of the journal Energy Sources. He is recognized as a world leader in the areas of fuels characterization and development. Dr. Speight is also Adjunct Professor of Chemical and Fuels Engineering at the University of Utah. James Speight is also a Consultant, Author and Lecturer on energy and environmental issues. He has a B.Sc. degree in Chemistry and a Ph.D. in Organic Chemistry, both from University of Manchester. James has worked for various corporations and research facilities including Exxon, Alberta Research Council and the University of Manchester. With more than 45 years of experience, he has authored more than 400 publications--including over 50 books--reports and presentations, taught more than 70 courses, and is the Editor on many journals including the Founding Editor of Petroleum Science and Technology.

Read more from James G. Speight

Related to Environmental Inorganic Chemistry for Engineers

Related ebooks

Agriculture For You

View More

Related articles

Related categories

Reviews for Environmental Inorganic Chemistry for Engineers

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Environmental Inorganic Chemistry for Engineers - James G. Speight

    Environmental Inorganic Chemistry for Engineers

    First Edition

    DR. James G. Speight

    CD & W inc., Laramie, Wyoming, United States

    Table of Contents

    Cover image

    Title page

    Copyright

    Author Biography

    Preface

    Chapter One: Inorganic Chemicals in the Environment

    Abstract

    1.1 Introduction

    1.2 The Environment

    1.3 Inorganic Chemistry and the Environment

    1.4 Use and Misuse of Inorganic Chemicals

    1.5 Inorganic Chemicals in the Environment

    1.6 Chemistry and Engineering

    Chapter Two: Inorganic Chemistry

    Abstract

    2.1 Introduction

    2.2 Nomenclature of Inorganic Compounds

    2.3 Classification of Inorganic Chemicals

    2.4 The Periodic Table

    2.5 Bonding and Molecular Structure

    2.6 Reactions and Stoichiometry

    2.7 Acid-Base Chemistry

    2.8 Minerals

    Chapter Three: Industrial Inorganic Chemistry

    Abstract

    3.1 Introduction

    3.2 The Inorganic Chemicals Industry

    3.3 The Production of Inorganic Chemicals

    3.4 Inorganic Polymers

    3.5 Inorganic Pigments

    Chapter Four: Properties of Inorganic Compounds

    Abstract

    4.1 Introduction

    4.2 Bond Lengths and Bond Strengths

    4.3 Physical Properties

    4.4 Critical Properties

    4.5 Reactive Chemicals

    4.6 Catalysts

    4.7 Corrosivity

    Chapter Five: Sources and Types of Inorganic Pollutants

    Abstract

    5.1 Introduction

    5.2 Sources

    5.3 Fly Ash and Bottom Ash

    5.4 Characterization of Inorganic Compounds

    Chapter Six: Introduction Into the Environment

    Abstract

    6.1 Introduction

    6.2 Minerals

    6.3 Release Into the Environment

    6.4 Types of Chemicals

    6.5 Physical Properties and Distribution in the Environment

    Chapter Seven: Transformation of Inorganic Chemicals in the Environment

    Abstract

    7.1 Introduction

    7.2 Chemical and Physical Transformation

    7.3 Inorganic Reactions

    7.4 Catalysts

    7.5 Sorption and Dilution

    7.6 Biodegradation

    7.7 Chemistry in the Environment

    Chapter Eight: Environmental Regulations

    Abstract

    8.1 Introduction

    8.2 Environmental Impact of Production Processes

    8.3 Environmental Regulations in the United States

    8.4 Outlook

    Chapter Nine: Removal of Inorganic Compounds From the Environment

    Abstract

    9.1 Introduction

    9.2 Cleanup

    9.3 Bioremediation

    9.4 Remediation of Heavy Metal-Contaminated Sites

    9.5 Pollution Prevention

    Appendix

    Conversion Factors

    1 Area

    2 Concentration Conversions

    3 Nutrient Conversion Factor

    4 Temperature Conversions

    5 Sludge Conversions

    6 Various Constants

    7 Volume Conversion

    8 Weight Conversion

    9 Other Approximations

    Glossary

    Index

    Copyright

    Butterworth-Heinemann is an imprint of Elsevier

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

    Copyright © 2017 Elsevier Inc. All rights reserved.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions.

    This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein).

    Notices

    Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary.

    Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility.

    To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein.

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    British Library Cataloguing-in-Publication Data

    A catalogue record for this book is available from the British Library

    ISBN: 978-0-12-849891-0

    For information on all Butterworth-Heinemann publications visit our website at https://www.elsevier.com/books-and-journals

    Publisher: Matthew Deans

    Acquisition Editor: Ken McCombs

    Editorial Project Manager: Peter Jardim

    Production Project Manager: Kiruthika Govindaraju

    Cover Designer: Vitoria Pearson

    Typeset by SPi Global, India

    Author Biography

    Dr. James G. Speight CChem., FRSC, FCIC, and FACS earned his BSc and PhD degrees from the University of Manchester, England; he also holds a DSc in geologic sciences and a PhD in petroleum engineering. Dr. Speight is the author of more than 70 books in fossil fuel and engineering and environmental sciences. Now, being an independent consultant, he was formerly the CEO of the Western Research and has served as adjunct professor in the Department of Chemical and Fuels Engineering at the University of Utah and in the Departments of Chemistry and Chemical and Petroleum Engineering at the University of Wyoming. In addition, he has also been a visiting professor in chemical engineering at the following universities: the University of Missouri-Columbia, the Technical University of Denmark, and the University of Trinidad and Tobago.

    Dr. Speight was elected to the Russian Academy of Sciences in 1996 and awarded the gold Medal of Honor that same year for the outstanding contributions to the field of petroleum sciences. He has also received the Scientists without Borders Medal of Honor of the Russian Academy of Sciences. In 2001, the academy also awarded Dr. Speight the Einstein medal for the outstanding contributions and service in the field of geologic sciences.

    Preface

    The latter part of the 20th century saw the realization arise that all chemicals can act as environmental pollutants, and in addition, there came the realization that emissions of inorganic chemicals such as carbon dioxide (CO2), methane (CH4), and nitrous oxide (N2O) to the atmosphere and a host of other inorganic chemicals to water systems and to land systems had either a direct or indirect impacts on the various floral and faunal systems. As a result, unprecedented efforts have been made to reduce all global emissions and chemicals disposal in order to maintain a green perspective. Furthermore, operations have been designed to reduce the direct emissions of inorganic chemical products (such as carbon dioxide, sulfur oxides, and nitrogen oxides) into the air (the atmosphere), inorganic chemicals into water systems (the aquasphere), and inorganic chemicals on to the soil (the terrestrial biosphere). This has been accompanied by efforts to recycle and to reuse as much of these chemicals and chemical waste as possible.

    The primary purpose of this book is to focus on the various issues related to the production and use of inorganic chemical issues that are the focus of any environmental chemistry program. In the context of inorganic chemistry, the book also presents an understanding of information on environmentally important physiochemical properties of inorganic chemicals for use by engineers and managers as well as any nonchemists. This book is a companion text to a book (Environmental Organic Chemistry for Engineers, James G. Speight, Elsevier, 2017) previously prepared for environmentally important effects and properties of organic chemicals. Like the organics book, this inorganics book describes available information that relates to the properties of inorganic chemicals and the effects of these chemicals on the environment. This information supports the primary objective of the book, which is to assist environmental engineers and managers (many of whom may not have a detailed knowledge or understanding of inorganic environmental chemistry) in overcoming the common problem of property data gaps and developing timely responses to environmental problems. This inorganics book presents not only generic discussion of these properties but also a summary of environmentally important data for the most common elements and pollutants.

    The topics covered in this book are the basis topics that serve to introduce the reader to not only inorganic chemistry but also the effect of inorganic chemicals on various ecosystems. Basic rules of nomenclature are presented. Understanding the mechanism of how a reaction takes place is particularly crucial in this and of necessity; the book brings a logic and simplicity to the reactions of the different functional groups. This in turn transforms a list of apparently unrelated facts into a sensible theme. Thus, this chapter will serve as an introduction to the physicochemical properties of inorganic chemicals and their effect on the floral and faunal environments.

    The book will serve as an information source the engineers in presenting details of the various aspects of inorganic chemicals as they pertain to pollution of the environment. To accomplish this goal, the initial section (Chapters 1–4) presents an introduction to and a description of the nomenclature of inorganic compounds and the properties of these materials. The remaining part of the book (Chapters 5–9) presents information relevant to the sources of inorganic contaminants, the behavior of inorganic chemicals, the fate and consequences of chemicals in the environment, and cleanup of the environment.

    From the book, the reader will gain an understanding of the fundamental inorganic chemistry and chemical processes that are central to a range of important environmental problems and to utilize this knowledge in making critical evaluations of these problems. Specific knowledge will be in the area of (i) an understanding of the chemistry of the stratospheric ozone layer and of the important ozone depletion processes; (ii) an understanding of the chemistry of important tropospheric processes, including photochemical smog and acid precipitation; (iii) understanding of the basic physics of the greenhouse effect and of the sources and sinks of the family of greenhouse gases; (iv) an understanding of the nature, reactivity, and environmental fates of toxic inorganic chemicals; and (v) an understanding of societal implications of some environmental problems. An appendix contains a selection of tables that contain data relating to the properties and characterization of the elements and inorganic compounds, and a comprehensive glossary will help the reader to understand the common terms that are employed in this field of science and engineering.

    Dr. James G. Speight, Laramie, Wyoming

    Jan. 2017

    Chapter One

    Inorganic Chemicals in the Environment

    Abstract

    Contamination by inorganic chemicals is a global issue, and such toxic chemicals are found practically in all ecosystems. It is the inappropriate management of such waste (e.g., through haphazard and unregulated disposal) that poses negative impacts on the environment. It is the purpose of this chapter to introduce the reader to the various aspects of inorganic chemicals and the environment and the role that inorganic can play dealing with the various issues of the environment. Throughout the pages of this book, the reader will be presented with the definitions and explanations of terms related to inorganic chemistry and how inorganic chemistry can be understood and used and the effects of the chemical on the environment.

    Keywords

    Environment; Atmosphere; Aquasphere; Terrestrial biosphere; Inorganic chemistry; Indigenous chemicals; Nonindigenous chemicals; Use and misuse of chemicals

    1.1 Introduction

    Environmental chemistry focuses on environmental concerns about materials, energy, and production cycles and demonstrates how fundamental chemical principles and methodologies can protect the floral (plant) and faunal (animal, including human) species within the environment (Anastas and Kirchhoff, 2002). More specifically, the principles of chemistry can be used to develop how global sustainability can be supported and maintained. For this, future environmental chemists and environmental engineers must acquire the scientific and technical knowledge to design products and chemical processes. They must also acquire an increased awareness of the environmental impact of chemicals on the environment and develop an enhanced awareness of the importance of sustainable strategies in chemical research and the chemical industry, specifically in the context of this book the inorganic chemicals industry.

    By way of introduction, although other classification systems have been published, a general classification of chemical pollutants is based on the chemical structure of the pollutant and includes (i) organic chemical pollutants and (ii) inorganic chemical pollutants. For the purposes of this text, organic chemical pollutants are those chemicals of organic origin or that could be produced by living organisms or are based of matter formed by living organisms (Speight, 2017a).

    On the other hand, inorganic chemical pollutants are those chemicals of mineral origin in (not produced by living organisms). In general, substances of mineral origin (such as ceramics, metals, synthetic plastics, and water) as opposed to those of biological or botanical origin (such as crude oil, coal, wood, and food). In addition, minerals are the inorganic, crystalline solid that makes up rocks. With certain exceptions, inorganic substances do not contain carbon or its compounds. In scientific terms, no clear line divides organic and inorganic chemistry.

    Inorganic chemistry focuses on the classification of inorganic compounds based on the properties of the compound(s) (Weller et al., 2014). Partly, the classification focuses on the position in the periodic table (Fig. 1.1) of the heaviest element (the element with the highest atomic weight) in the compound, partly by grouping compounds on the basis of structural similarities. Also, inorganic compounds are generally structured by ionic bonds and do not contain carbon chemically bound to hydrogen (hydrocarbons) or any of their derivatives that contain elements such as nitrogen, oxygen, sulfur, and metals. Examples of inorganic compounds include sodium chloride (NaCl) and calcium carbonate (CaCO3) and pure elements (Cox, 1995).

    Fig. 1.1 The periodic table of the elements.

    Thus, this text relates to an introduction to the planned and unplanned effects of inorganic chemicals on the various environmental systems. Inorganic chemicals are an essential component of life, but some chemicals are extremely toxic and can severely damage the floral (plant life) and faunal (animal life) environment (Table 1.1).

    Table 1.1

    Examples of the Classification of Elements According to Their Effects in the Biological Systemsa

    Essential

    O, C, H, N, P, Na, K, Mg, Cl, Ca, S

    Trace

    I, Fe, Cu, Zn, Mn, Co, Mo

    Nonessential

    Al, Sr, Ba, Sn

    Toxic

    Cd, Pb, Hg

    a Essential elements are necessary for life processes; trace elements are also necessary for life processes; nonessential elements are not essential. If they are absent, other elements may serve the same function; toxic elements disturb the natural functions of the biological system.

    As with organic chemicals (Speight, 2017a), contamination of the environment by inorganic chemicals is a global issue, and toxic inorganic chemicals are found practically in all ecosystems because, at the end of the various inorganic chemical life cycles, inorganic chemicals have been either recycled for further use or sent for disposal as chemical waste (Bodek et al., 1988). Current regulations do not permit the unmanaged disposal of chemical waste but, in the past (particularly in the first decades of the 20th century), the inappropriate management of chemical waste (e.g., through haphazard disposal and unregulated burning) has led to a series of negative and lingering impacts on the floral and faunal species that are part of the environment.

    H bonds). The distinction between the two subdisciplines is far from absolute, as there is much overlap within the subdiscipline of organometallic chemistry. Nevertheless, the principles of inorganic chemistry have application in every aspect of the chemical industry, including materials science, catalysis, surfactants, pigments, coatings, medications, fuels, and agriculture.

    Many inorganic compounds are ionic compounds, consisting of cations and anions joined by ionic bonding (ions in ammonium carbonate. In any salt, the proportions of the ions are such that the electric charges cancel out, so that the bulk compound is electrically neutral. The ions are described by their oxidation state, and their ease of formation can be inferred from the ionization potential (for cations) or from the electron affinity (anions) of the parent elements (Chapter 2). Ionic compounds containing hydrogen ions (H+) are classified as acids, and those containing basic ions, such as the hydroxide anion (OH−) or oxide anion (O²−), are classified as bases. Some ions are classed as amphoteric because of the ability of these ions to react with either an acid or a base (Davidson, 1955). This is also true of some compounds with ionic character, typically oxide derivatives or hydroxide derivatives of the less electropositive metals (which results in the compound having significant covalent character), such as zinc oxide (ZnO) aluminum hydroxide [Al(OH)3], aluminum oxide (Al2O3), and lead(II) oxide (PbO).

    Cl−1)—all of these derivatives occur as minerals in the Earth (Chapter 2). Many inorganic compounds are characterized by high melting points (Chapter 4). Other important features include their high melting point and ease of crystallization, where some salts are very soluble in water (such as sodium chloride, NaCl) and crystallize form concentrated solutions of the salt others (such as silica—silicon dioxide, SiO2). Thus,

    Inorganic chemistry: oxygen and metals are the dominant elements.

    Organic chemistry: carbon and hydrogen are the dominant elements.

    By way of further definition and to alleviate any potential confusion, bioinorganic chemistry (bioinorganic chemistry and biological inorganic chemistry) is a subcategory of chemistry that examines the role of metals in biology systems including the associated environmental issues (Bertini et al., 1994; Fraústo da Silva and Williams, 2001). Thus, bioinorganic chemistry includes the study of both natural phenomena such as the behavior of metalloproteins and artificially introduced metals, including those that are nonessential (Table 1.1) in medicine and toxicology. Many biological processes, such as respiration, depend upon the molecular species that fall within the realm of inorganic chemistry and, as a blend of biochemistry and inorganic chemistry, bioinorganic chemistry is important in elucidating the activity of proteins and the effect of the properties of inorganic compounds as they pertain to the activity and well-being of floral and faunal organisms and an understanding of the various elements inorganic compounds on floral and faunal systems.

    As a side but relevant note, it must be assumed that all chemicals are toxic unless proved otherwise. In relation to the environmental effects of inorganic chemicals, consideration must also be given to the effect of the so-called harmless chemicals (the indigenous chemicals and the natural products chemicals) when they are added back to the environment in amounts that exceed the natural abundance. Within the local environment, these chemicals will be present in a measurable concentration, but the flora and fauna present in that ecosystem may be fatally susceptible to the effects of such chemicals when they are present in a concentration that is above the indigenous (natural) concentration of the chemicals. Using the human experience as an example, a sprinkling salt on a cooked vegetable or meat may add to the taste of the meal. However, it is extremely dangerous (the consequences can even be fatal), and therefore, unadvisable for a human to attempt to consume several ounces of salt with that same meal. Not only would the taste be ruined, but also the high concentration of salt could have a serious health effects (even death) on the consumer.

    Thus, there is an increase in environmental issues and problems that can be partially explained because of the use of chemicals. In fact, many man-made chemicals are found in the most remote places in the environment. However, in order to successful manage the environment and protect the flora and fauna from such chemicals, knowledge of chemical, specifically inorganic chemicals (in the context of this book), is a decided advantage. Moreover, the chief reason for studying this subject is not only to understand the effects of inorganic chemicals on the environment, which may be caused by unforeseen side effects of a chemical substance during its production, transport, use, and disposal. These effects provide the motivation for the build of scientific knowledge on the effects of inorganic chemicals on the floral and faunal environments. Ideally, environmental scientists and environmental engineers should be able to predict the possible (if not, likely) effects of an inorganic chemical directly on the environment even before the chemical substance is released, enabling a more realistic appraisal to be made of any effects of the chemical on an ecosystem.

    Thus, a first approximation to predicting a potentially harmful inorganic chemical may involve the following criteria: (i) whether the chemical is biologically essential; (ii) whether the chemical is biologically nonessential; (iii) whether the chemical is toxic in a variety of concentrations; (iv) whether the chemical is likely to undergo transformation in the environment and form highly stable inert compounds; (v) whether the chemical is unlikely to undergo transformation in the environment and form highly stable inert compounds; (vi) whether the chemical persists in the environment; (vii) whether the chemical interacts favorably or adversely with the flora and fauna of the ecosystem into which the chemical is released; (viii) whether the chemical is environmentally mobile in, or interferes with, any of the biological or biogeochemical cycles; and (ix) whether the chemical is environmentally immobile in, or interferes with, any of the biological or biogeochemical cycles.

    Thus, as for any technical discipline, the nonchemist is faced with understanding many new terms that are related to the discipline of inorganic chemistry and that need to be understood to place inorganic chemicals in the correct environmental context. Some terms may seem familiar, but in chemistry, especially in inorganic chemistry where the terms may have meanings that are not quite the same as when used in popular commentaries, even at various technical meetings. For example, in the subdiscipline of inorganic chemistry and, indeed, in all subdisciplines of chemistry, the terms need to have definite and specific meanings to make the subject matter understandable, and this is the raison d'être for the acceptable systems of nomenclature (Chapter 2). One of the purposes of any system of chemical nomenclature is to provide definitions for many of these terms in a form and at a level that will make the meaning clear to technical persons in other discipline who need to deal with the various aspects of chemistry.

    In this respect, the International Union of Pure and Applied Chemistry (IUPAC)—an international organization of chemists and national chemistry societies and the world authority on chemical nomenclature and terminology with a secretariat in Research Triangle Park, North Carolina, https://iupac.org/who-we-are/—makes the final determination of terminology and nomenclature in chemistry. Among other projects, this organization (IUPAC) authorizes and establishes systematic rules for naming compounds so that any inorganic (or organic) chemical structure can be defined uniquely and in a meaningful manner. Compounds are frequently called by common names or by trade names, often because the IUPAC names may be long and complex and difficult to understand for the nonchemist or the company need to protect the identity of the chemical; this is especially true in the pharmaceutical industry where the true (correct) chemical name could reveal the company secrets of drug synthesis. However, the IUPAC name permits the chemist and the nonchemist to know the structure of any chemical compound based on the rules of the IUPAC terminology, while the common name or trade name may be outside of the realms of the reality of chemical nomenclature.

    Engineering students, on the other hand, undergo a much different form of training and education when compared with training to which the student of chemistry is subjected. While some (but not all) engineering disciplines (especially the chemical engineer) may require some background knowledge of chemistry, the practicing engineer is more concerned with practical applications of chemical synthesis (such as reactor construction, reactor operation, process parameters, and product yield), and there are differences in, for example, reactor novelty and reactor scale—areas that are not always pertinent to the chemist outside of laboratory chemistry who is accustomed to employing scientific glassware for his/her laboratory experiments.

    Thus, a chemist is more likely to be engaged in developing new compounds and materials using novel or new synthetic routes, while a chemical engineer is more likely to be working with existing substances and improving the reaction parameters to suit industrial synthesis of the chemical. A chemist may be involved in laboratory synthesis program to produce several grams of a new compound, while a chemical engineer will focus on scale up of the synthetic process to produce the chemical (say, on a tonnage scale) at a profit. Thus, the chemical engineer will be more concerned with heating and cooling large reaction vessels, pumps and piping to transfer chemicals, plant design, plant operation, and process optimization, while a chemist will be more concerned with establishing the parameters of the reaction from which the engineer will design the plant. However, these differences are generalizations, and there is often much overlap. As a consequence, and throughout the pages of this book, the reader will be presented with the definitions and explanations of terms related to inorganic chemistry and the means by which inorganic chemistry can be understood and used (Weller et al., 2014).

    An alternative quantitative approach to inorganic chemistry focuses on properties of inorganic chemicals and the energies of the various reactions (Chapter 4). This approach is highly traditional and empirical, but it is useful in terms of understanding the behavior of inorganic chemicals in the environment (Chapter 7). The general concepts of inorganic chemistry that fall under the umbrella of thermodynamics include (i) acidity or alkalinity, (ii) redox potential, and (iii) phase changes in addition to (iv) the Born-Haber cycle, which is used for assessing the energies of elementary processes such as electron affinity and hence theoretical reactivity of inorganic chemicals (Moore and Stanitski, 2014).

    1.2 The Environment

    There is the need for two definitions to enhance the understanding of the concepts presented in this book, thus: (i) ecosystem and (ii) environment. There is a general tendency to use these terms interchangeably, but to be more correct, the definitions are (i) the term ecosystem relates to a community of organisms together within their physical environment, viewed as a system of interacting and interdependent relationships and including such processes as the flow of energy through trophic levels and the cycling of chemical elements and compounds through living and nonliving components of the system and (ii) the term environment relates to the conditions that surround someone or something; the conditions and influences that affect the growth, health, progress, etc., of someone or something; the total living and nonliving conditions (internal and external surroundings) that are an influence on the existence and complete life span of the organism.

    Thus, the natural environment encompasses all living and nonliving flora and fauna that occur naturally and environment encompasses the interaction of all living species, climate, weather, and natural resources that affect survival and activity of the flora and fauna (Manahan, 2009). Furthermore, the concept of the natural environment can be distinguished by components: (i) complete ecological units that function as natural systems without massive civilized human intervention, including the atmosphere, all vegetation, microorganisms, soil, rocks, and natural phenomena that occur within the boundaries of the ecological unit and (ii) the natural resources and physical phenomena that lack clear-cut boundaries, such as air, water, and climate, as well as energy, radiation, and other effects that do not originate from anthropogenic (human) activity. In contrast to the natural environment is the anthropogenic environment in which human activity has fundamentally transformed landscapes such as urban settings and agricultural land conversion, the natural environment is greatly modified into a simplified human environment.

    In fact, there has been the suggestion (from the International Geological Congress in Cape Town) that the name of the present geologic name of the Holocene (Fig. 1.2) should be changed to Anthropocene (Pöschl and Shiraiwa, 2015; Robinson, 2016). This is an epoch that is defined by human impact on the environment and is an indication that humans have changed the planet. However, the proponents of this change—while correct to a point—seem to have ignored the fact that the Earth is in an interglacial period when the climate is expected to change (as the Earth warms due to natural effects), and the human effects may be somewhat less than calculated (Speight and Islam, 2016). Defining an epoch normally relies on evidence laid down over thousands or millions of years and whether renaming the Holocene to the Anthropocene remains to be clarified and fully justified.

    Fig. 1.2 Geologic era, epochs, and period.

    Renaming the Holocene aside, to understand the influence of any chemical on the environment, it is first necessary to understand the structure of the environment though an understanding of the and the various subdivisions of the air, water, and land. When an inorganic chemical or any chemical for that matter (Speight, 2017a) is introduced into the environment, it is eventually distributed among the four major environmental compartments, which are often referred to as separate but interrelated ecosystems: (i) air, (ii) water, (iii) soil, and (iv) biota (living organisms). Depending upon the situation, further subdivision of the categories can be made into (i) floral systems, i.e., plant systems and (ii) faunal systems, i.e., animals, including human systems.

    The fraction of the chemical that will move into each environmental subdivision is governed by the physical properties and the chemical properties of the chemical (Table 1.2) (Hickey, 1999; Yaron et al., 2011). In addition, the distribution of inorganic chemicals in the environment is governed by physical processes such as solubility in aqueous systems, sedimentation, adsorption, and volatilization evaporation, or sublimation and the mechanisms by which these chemicals can be degraded by chemical processes and/or biological processes. In contrast, the solubility of organic chemicals in aqueous systems is somewhat less than the solubility of inorganic chemicals, whereas organic chemicals (Speight, 2017a) are typically less prone to being soluble in aqueous systems but may be more subject to evaporation or vaporization than inorganic chemicals. In addition, chemical transformation processes (Chapter 7) generally occur in the atmosphere, in water, or in the soil and follow one of four reactions such as oxidation, reduction, hydrolysis, and photolysis.

    Table 1.2

    Important Common Physical Properties of Inorganic Compounds

    State: Gas

    Density

    Critical temperature, critical pressure (for liquefaction)

    Solubility in water, selected solvents

    Odor threshold

    Color

    Diffusion coefficient

    State: Liquid

    Vapor pressure-temperature relationship

    Density; specific gravity

    Viscosity

    Miscibility with water, selected solvents

    Odor

    Color

    Coefficient of thermal expansion

    Interfacial tension

    State: Solid

    Melting point

    Density

    Odor

    Solubility in water, selected solvents

    Coefficient of thermal expansion

    Hardness/flexibility

    Particle size distribution/physical form, such as fine powder, granules, pellets, lumps

    Porosity

    Biological mechanisms in soil and living organisms typically utilize oxidation, reduction, and hydrolysis to degrade chemicals (Speight and Arjoon, 2012). The process of degradation will largely be governed by the compartment (water, soil, atmosphere, and biota) in which the inorganic chemical is distributed and the distribution is governed, in turn, by the physical processes already mentioned (i.e., solubilization, sedimentation, adsorption, and volatilization/evaporation/sublimation). The latter processes—i.e., volatilization, evaporation, and sublimation—are generally less obvious or effective in the field of inorganic chemistry as they are in the field of organic chemistry (Speight, 2017a).

    The impact on the environment of the transformation gin the chemical state of inorganic chemicals is only partially elucidated but can be expected to be significant in many cases. Changes in the atmospheric abundance of various gases have been claimed (with some justification) to lead to observable changes in the climate of the Earth, including (i) changes in temperature, (ii) patterns of precipitation, and (iii) the frequency of occurrence of extreme natural events, such as hurricanes. Ecosystem damage also results from regional and global pollution. Moreover, acidic precipitation (acid rain) is claimed to be the reason for the suppression life in several lakes of North America and Europe and, together with enhanced ozone (O3) levels, to have damaged forests in those same parts of the world.

    In fact, when assessing the impact of inorganic chemicals on the environment, the most critical characteristics are (i) the types and properties of the discharged chemicals, which depend on the type of chemicals and processes used to produce the chemicals and (ii) the amount and concentration of the discharged chemicals. Solid waste (containing inorganic chemicals) and/or gaseous emissions generated from industrial sources also contribute to the amount and concentration (and effects) of inorganic chemicals in the environment.

    This has led to the introduction of the concept of an emissions factor that is a representative value that attempts to relate the quantity of a pollutant released to the atmosphere with an activity associated with the release of that pollutant. The emission factor is usually expressed as the weight of pollutant divided by a unit weight, volume, distance, or duration of the activity emitting the pollutant (such as the kilogram of pollutant per kilogram of produced product). The factor facilitates an estimation of emissions from various sources of air pollution. In most cases, these factors may be averages of all available data of acceptable quality, which are generally (correctly or incorrectly) assumed to be representative of long-term averages for all facilities in the source category (i.e., a population average). Thus, the general equation for the estimation of emissions is

    In this equation, E=emissions, A=activity rate, EF=emission factor, and ER=overall emission reduction efficiency, %.

    However, caution is advised when making such assumptions because the word average is not always indicative of the damage to the environment by that amount of chemical. The upper level data (used to compute the average) can be amounts of the chemical that lead to severe destruction of an ecosystem. On the other hand, the lower level data (used to compute the average) can be amounts of the chemical that can be tolerated by the ecosystem without causing serious disruption of that ecosystem. In either case, if there are data points that are flyers (data points that lie outside of the acceptable limits of variance—another assumption that may be erroneous), such data may skew the factor to the high end or low end. Or, if there are data points that could be classed as flyers and that have been included in the calculation of the factor, such data may skew the factor to the high end or low end.

    1.2.1 Structure of the Atmosphere

    The atmosphere is the gaseous envelope that surrounds the solid body of the Earth. Although the atmosphere has a thickness on the order of 600 miles, approximately half of the atmospheric mass is concentrated in the lower 3 miles. The atmosphere (i) is a protective blanket that nurtures life on the Earth and protects the planet from the hostile environment of outer space; (ii) is the source of carbon dioxide for plant photosynthesis and of oxygen for respiration; (iii) provides the nitrogen that nitrogen-fixing bacteria and ammonia-manufacturing industrial plants use to produce chemically bound nitrogen, an essential component of molecules of life; (iv) transports water from the oceans to land, thus acting as the condenser in a vast solar powered still; and (v) serves a vital protective function, absorbing harmful ultraviolet radiation from the sun and stabilizing the temperature of the Earth.

    Physically, the atmosphere of the Earth is the thin and fragile envelope of air surrounding the Earth that is held in place around the Earth by gravitational attraction and that has a substantial effect on the environment. The total dry mass of the atmosphere (annual mean), three quarters of which is within approximately 36,000 ft of the surface, is estimated to be more than 5×10²¹ tons (Trenberth and Guillemot, 1994). In addition, the atmosphere is largely transparent and allows incoming sunlight to reach the surface of the Earth, provided the sunlight is not reflected or absorbed by clouds.

    A fraction of the light that does reach the planet is absorbed, based on the degree of reflectivity of the surface, and most of the absorbed light is converted to energy, and heat is radiated back out at infrared wavelengths. Although the atmosphere allows most of the visible light through, many of the inorganic gases, such as water vapor (H2O) and carbon dioxide (CO2) absorb infrared radiation, thereby converting the infrared radiation to rotational and vibrational energy. This raises the energy content of the atmosphere and the average temperature. Thus, the higher the content of these gases in the atmosphere, the greater the chance of the infrared light being absorbed before it escapes into space.

    If all other influences are maintained in a constant state, increased levels of greenhouse gases will necessarily produce increased atmospheric temperatures. However, the impact of greenhouse gases (any of the gases whose absorption of solar radiation is responsible for the greenhouse effect, including carbon dioxide, ozone, methane, and the fluorocarbons) differs based on the chemistry of the individual gases. Moreover, the impact of any gas (greenhouse or otherwise) is influenced by the lifetime of the gas in the atmosphere; for example, water vapor is removed from the atmosphere as precipitation, while organic gases (not the subject of this book but worthy of mention here) such as methane are typically oxidized to carbon dioxide and water within decades of its appearance in the atmosphere:

    In general, air pressure and air density decrease with altitude in the atmosphere, but the influence of the temperature (which has a more complicated profile with altitude) may remain relatively constant or even increase with altitude in some regions. Because the general pattern of the temperature-altitude profile is relatively constant and measurable (by means of instrumented balloon soundings), the temperature behavior provides a useful paradigm by which the atmospheric layers can be subcategorized.

    Thus, the atmosphere can be divided (called atmospheric stratification) into five primary areas that differ in properties such as composition, temperature, and pressure and that include, from the lowest to the highest, the five main layers: (i) the troposphere, (ii) the stratosphere, (iii) the mesosphere, (iv) the thermosphere, and (v) the exosphere. The exosphere is the highest but thin, atmosphere-like volume surrounding the Earth that is gravitationally bound to the Earth, but the density is too low for any molecular collisions to occur on a regular basis. Although, in a more general sense, the atmosphere is defined by the homosphere and the heterosphere, in turn, they are defined by whether the atmospheric gases are well mixed.

    By way of explanation, the surface-based homosphere includes the troposphere, stratosphere, mesosphere, and the lowest part of the thermosphere, where the chemical composition of the atmosphere is not fully dependent on the molecular weight of the gaseous constituents because the gases are mixed by turbulence. This relatively homogeneous layer ends at the turbopause found at approximately 62 miles (330,000 ft), which places it approximately 12 miles (66,000 ft) above the mesopause.

    The heterosphere lies above the homosphere and includes the exosphere and most of the thermosphere. In this atmospheric layer, the chemical composition varies with altitude because the distance that particles can move without colliding with one another is large compared with the extent of the size of motion that causes mixing. This allows the gases to stratify (segregate) by molecular weight, with gases such as oxygen and nitrogen present only near the bottom of the heterosphere. The upper part of the heterosphere is composed almost completely of hydrogen, the lightest element.

    The planetary boundary layer is the part of the troposphere that is closest to the surface of the Earth and is directly affected by it, mainly through turbulent diffusion. During the day, the planetary boundary layer usually is well mixed, whereas at night it becomes stably stratified with weak or intermittent mixing. The depth of the planetary boundary layer ranges from as little as approximately 300 ft on clear, calm nights to 10,000 ft or more during the afternoon in dry regions.

    Approximately three quarters (75%, v/v) of the atmosphere's mass resides within the troposphere and is the layer within which the weather systems develop. The depth of this layer varies between 548,000 ft at the equator to 23,000 ft over the polar regions. The stratosphere, extends from the top of the troposphere to the bottom of the mesosphere, contains the ozone layer that ranges in altitude between 49,000 and 115,000 ft, and is where most of the ultraviolet radiation from the Sun is absorbed. The top of the mesosphere, ranges from 164,000 to 279,000 ft and is the layer wherein most meteors burn up. The thermosphere extends from 279,000 ft to the base of the exosphere at approximately 2,300,000 ft altitude and contains the ionosphere, a region where the atmosphere is ionized by incoming solar radiation.

    1.2.1.1 Composition

    Thus, the atmosphere is a mixture of chemical constituents—the most abundant of are nitrogen (N2, (78%, v/v)) and oxygen (O2, (21%, v/v)). These gases and the noble gases (argon, neon, helium, krypton, and xenon) possess very long lifetimes against chemical destruction and hence are relatively well mixed throughout the entire homosphere. Minor constituents, such as water vapor, carbon dioxide, ozone, and many others, also play an important role despite their lower concentration. These constituents influence the transmission of solar radiation and terrestrial radiation in the atmosphere and are therefore linked to the physical climate system. They are also key components of biological cycles and biogeochemical cycles and, in addition, they determine the oxidizing capacity of the atmosphere and hence the atmospheric lifetime of biogenic gases (gases produced by life processes) and anthropogenic gases (gases produced by human activities).

    Water vapor accounts for approximately 0.25% (v/v) (~0.25%, w/w) of the atmosphere (Table 1.3), but the concentration of water vapor (a greenhouse gas) varies significantly from approximately 10 ppm (v/v) in the coldest portions of the atmosphere to as much as 5% (v/v) in hot, humid air masses. The concentrations of other atmospheric gases are typically quoted in terms of dry air (without water vapor) and are often referred to as trace gases, among which are the inorganic greenhouse gases, principally carbon dioxide (CO2), nitrous oxide (N2O), and ozone (O3). The spatial and temporal distribution of chemical species in the atmosphere is determined by several processes that include (i) surface emissions, (ii) deposition, (iii) chemical reactions, (iv) photochemical reactions, and (v) transport. Surface emissions are associated with volcanic eruptions; floral and faunal activity on the continental land masses and in the oceanic activities; and anthropological activity such as biomass burning, agricultural practices, and industrial activity. Wet deposition results from precipitation of soluble species within the moisture, while the rate of dry deposition is affected by the nature of the surface (such as the type of soil, the type of vegetation, and the presence of a large water system, such as an ocean).

    Table 1.3

    Composition of the Atmosphere

    a Water vapor is not included in above dry atmosphere and is approximately 0.25% (v/v) over the full atmosphere but does vary considerably.

    Furthermore, knowledge of the inorganic components of biomass feedstock is important for process control and for handling coproducts and wastes resulting from energy and fuel utilization of biomass. Analytic survey of forestry thinnings (wood chips), agricultural residues (rice straw, wheat straw, and corn stover), and dedicated perennial grass crops (switchgrass, wheatgrass, and miscanthus) shows that, potentially, the whole periodic table may be present in biomass. The main effect of ashing is bonding of oxygen in the ash mainly as silicate, oxides, hydroxides, phosphates, and carbonate residual minerals. Carbon is partially retained as carbonates and graphite (char). Nitrogen is dominantly released to the flue gas, while sulfur is mostly retained in the ash as sulfates. Small losses (~19%, w/w) for both sulfur and chlorine were detected during ashing at 575°C (1070°F). The majority of the alkali metals (lithium, Li; sodium, Na; potassium, K; and rubidium, Rb) will substantially modify soil if applied as a fertilizer. Only magnesium (Mg), calcium (Ca), and strontium (Sr) of the alkali earth metals; manganese (Mn), copper (Cu), and zinc (Zn) of the period 4 transition metals; and molybdenum (Mo) and cadmium (Cd) of the period 5 transition metals may exceed regulatory limits if used as a fertilizer. The heavy elements occur in concentrations too low to cause concern except for selenium, Se. The high alkali content of some biomass ash thus makes them good candidates for use as fertilizers provided that they are applied in low proportions (<50%) to soil. Ash of wood material is a carrier of many of the alkali elements (Li, K, Rb, Mg, Ca, Sr, and Ba) and some transition elements (Mn, Cu, Zn, Mo, Ag, and Cd). In contrast, ash of herbaceous plant material is in addition to K only variably enriched: Li, Na, Se, and Mo in wheatgrass; Mg and Ca in switchgrass; Mg, Ca, and Cd in corn stover; Mn in rice straw; Mo in wheat straw; and Mn and Cd in miscanthus. Water leaching results in significant losses for anionic chlorine and sulfur and for most of the alkali metals, thus making resulting ash from such treated feedstock less attractive as potassium-containing fertilizers although fuel properties are enhanced for thermal conversion (Thy et al., 2013).

    By volume, dry air contains (subject to minor rounding of the data) 78.09% nitrogen, 20.95% oxygen, 0.93% argon, 0.039% carbon dioxide, and small amounts of other gases (Table 1.3). Air also contains a variable amount of water vapor, approximately 1% (v/v) at sea level and 0.4% (v/v) throughout the entire atmosphere. Air content and atmospheric pressure vary at different layers, and air suitable for use in photosynthesis by terrestrial plants and breathing of terrestrial animals is found only in the troposphere.

    1.2.1.2 Chemical Activity

    Some examples of atmospheric pollutants include nitrogen dioxide (NO2), sulfur dioxide (SO2), and carbon monoxide (CO). Furthermore, nitrogen dioxide and sulfur dioxide combine with water to form acids:

    These acids contribute to the long-term destruction of the environment due to the generation of acid rain. Carbon monoxide, generated by the incomplete combustion of hydrocarbons fuels and other carbonaceous materials, displaces and prevents oxygen from binding to hemoglobin and causes asphyxiation. Also, carbon monoxide binds with metallic pollutants and causes them to be more mobile in air and water. These reactions greatly reduce the protective effects of ozone against ultraviolet radiation.

    Chemical compounds released at the surface by natural and anthropogenic processes are oxidized in the atmosphere before being removed by wet or dry deposition. Key chemical species of the troposphere include oxygenated inorganic species and carbon monoxide and nitrogen oxides (nitric oxide, NO; nitrous oxide, N2O; nitrogen dioxide, NO2; and dinitrogen tetroxide, N2O4, which forms an equilibrium mixture with nitrogen dioxide):

    Nitrogen dioxide is also produced by lightning discharges in thunderstorms and nitric acid (HNO3) and peroxyacetyl nitrate (PAN, an unstable secondary pollutant present in photochemical smog and that decomposes into peroxyethanoyl radicals and nitrogen dioxide).

    Other chemical species include hydrogen compounds—specifically the hydroxy radical, OH, and the hydroperoxy radical, HO2, as well as hydrogen peroxide, H2O2; ozone, O3; and sulfur compounds (dimethyl sulfide (DMS) CH3SCH3; sulfur dioxide, SO2; and sulfuric acid, H2SO4). The hydroxyl radical (OH) has the capability of reacting with and efficiently destroying a large number of inorganic chemical compounds and hence has the potential to contribute directly to the oxidative capacity (reactivity) of the atmosphere. Ozone also plays an important role in the troposphere: together with water vapor ozone, it is the source of the hydroxy radical, and in addition, it contributes to climate forcing.

    By way of explanation, climate forcing is any influence on climate that originates from outside the climate system itself and is a major cause of climate change. This includes the temperature rise of the Earth during an interglacial period that exists at the present. The presence of ozone in the troposphere results not only from the intrusion of ozone-rich stratospheric air masses through the tropopause but also from photochemical reactions involving hydrocarbon derivatives, nitrogen oxides (NOx), and carbon monoxide (CO). One major question is to what extent the oxidizing capacity of the atmosphere has changed because of human activities.

    Finally, the release of sulfur-containing chemicals at the surface of the Earth and the subsequent oxidation of the sulfur compounds in the atmosphere leads to the formation of small liquid or solid particles that remain in suspension in the atmosphere. The release to the atmosphere of sulfur compounds is a result of human activities, specifically coal combustion and the combustion of various petroleum products (Speight and Arjoon, 2012; Speight, 2013, 2014).

    1.2.2 The Aquasphere

    The aquasphere (also called the hydrosphere and, on occasion also referred to as the aquatic biome) is the layer of water that, in the form of the oceans, rivers, and lakes, covers approximately 71% of the surface of the Earth (some estimate put this at 75%); more than 97% (v/v) of this water exists in oceans (Charette and Smith, 2010). The oceans present a vast reservoir of saltwater, on land as surface water in lakes and rivers, underground as groundwater, in the atmosphere as water vapor, in the polar icecaps as solid ice, and in many segments of the anthrosphere such as in boilers or municipal water distribution systems.

    Water is an essential part of all living systems (Jackson et al., 2001) and is the medium from which life evolved and in which life exists. Water also (i) carries energy and matter are through various spheres of the environment, (ii) leaches soluble constituents from mineral matter and carries them to the ocean or leaves them as mineral deposits some distance from their sources, (iii) carries plant nutrients from soil into the bodies of plants by way of plant roots, and (iv) absorbs solar energy in oceans, and this energy is carried as latent heat and released inland when it evaporates from oceans; the accompanying release of latent heat provides a large fraction of the energy that is transported from equatorial regions of the Earth to the poles of the Earth and is the energy behind storms.

    Fresh, clean, and drinkable water is a necessary but limited resource. Industrial, agricultural, and domestic wastes can contribute to the pollution of this valuable resource, and water pollutants can damage human and animal health. Three important classes of water pollutants are (i) inorganic pollutants, of which heavy metals are a part, and organic pollutants (Speight, 2017a). Heavy metals include transition metals such as cadmium, mercury, and lead, all of which can contribute to pollution and have serious effects on the floral species (plants) and faunal species (animals) in the environment. Inorganic pollutants such as hydrochloric acid (HCl), sodium chloride (NaCl), and sodium carbonate (NaCO3) change the acidity or alkalinity (the pH) and the salinity of the water, making it undrinkable or unsuitable for the support of animal and plant life. These effects can result in dire consequences for higher mammals such as humans.

    Due to overuse, pollution, and ecosystem degradation, the sources of most freshwater supplies—groundwater (water located below the soil surface), reservoirs, and rivers—are under severe and increasing environmental stress. The majority of the urban sewage in developing countries is discharged untreated into surface waters such as rivers and harbors. Approximately 65% (v/v) of the global freshwater supply is used in agriculture, and 25% (v/v) is used for industrial processes. Freshwater conservation therefore requires a reduction in wasteful practices such as inefficient irrigation, reforms in agriculture and industry, and strict pollution controls worldwide.

    Aquatic regions house numerous species of plants and animals, both large and small. In fact, this is where life began billions of years ago when amino acids first started to come together to form the more complex proteins. Without water, most life forms would be unable to sustain themselves, and the Earth would be a barren, desert-like place. Although water temperatures can vary widely, aquatic areas tend to be more humid and the air temperature on the cooler side. The aquasphere can be broken down into two basic regions: (i) freshwater regions, such as ponds and rivers, and (ii) marine regions, i.e., the oceans.

    1.2.2.1 Freshwater Regions

    A freshwater region is an area where the water has a low salinity (a low salt concentration, usually on the order of <1%, w/w). Plants and animals in freshwater regions are adjusted to the low salt content and would not be able to survive in areas of high salt concentration (such as the ocean). There are different types of freshwater regions: (i) ponds and lakes, (ii) streams and rivers, and (iii) wetlands (Speight, 2017a).

    Pond and lakes range in size from just a few square meters to thousands of square kilometers. Many ponds are seasonal, lasting just a couple of months (such as sessile pools), while lakes may exist for hundreds of years or more. Ponds and lakes may have limited species diversity since they are often isolated from one another and from other water sources like rivers and oceans. The temperature varies in ponds and lakes seasonally. During the summer, the temperature can range from 4°C (39°F) near the bottom to 22°C (72°F) at the top. During the winter, the temperature at the bottom can be 4°C (39 F) while the top is 0°C (32°F, ice). In between the two layers, there is a narrow zone called the thermocline where the temperature of the water changes rapidly. During the spring and fall seasons, there is a mixing of the top and bottom layers, usually due to winds, which results in a uniform water temperature of around 4°C (39°F). This mixing also circulates oxygen throughout the lake. There are many lakes and ponds that do not freeze during the winter, in which case the top layer would be a little warmer.

    Streams and rivers are bodies of flowing water moving in one direction; they typically start at headwaters, which may be springs, snowmelt or even lakes, and then travel all the way to their mouths, usually another water channel or the ocean. The characteristics of a river or stream change during the journey from the source to the mouth. The temperature is cooler at the source than it is at the mouth. The water is also clearer and has higher oxygen levels, and freshwater fish such as trout and heterotrophs can be found there. Toward the middle part of the stream/river, the width increases, as does species diversity; numerous aquatic green plants and algae can be found. Toward the mouth of the river/stream, the water becomes murky from all the sediments that it has picked up upstream, decreasing the amount of light that can penetrate through the water. Since there is less light, there is less diversity of flora, and because of the lower oxygen levels, fish that require less oxygen, such as catfish and carp, can be found.

    Wetlands are areas of standing water that support aquatic plants. Marshes, swamps, and bogs are all considered wetlands. Plant species adapted to the very moist and humid conditions are called hydrophytes. These include pond lilies, cattails, sedges, tamarack, and black spruce. Marsh flora also includes such species as cypress and gum. Wetlands have the highest species diversity of all ecosystems. Many species of amphibians, reptiles, birds (such as ducks and waders), and furbearers can be found in the wetlands. Wetlands are not considered freshwater ecosystems as there are some, such as salt marshes, that have high salt concentrations, which support different species of animals, such as shrimp, shellfish, and various grasses.

    1.2.2.2 Marine Regions

    Marine regions, more often referred to as oceans, cover approximately 71% of the surface of the Earth and that also (under the general term marine regions) includes coral reefs and estuaries; estuaries are areas where freshwater streams or rivers merge with the ocean (Charette and Smith, 2010). This mixing of waters with such different salt concentrations creates a very interesting and unique ecosystem. Microflora (such as algae) and macroflora (such as seaweeds, marsh grasses, and mangrove trees) can be found in these regions and, in fact, estuaries support a diverse fauna, including a variety of worms, oysters, crabs, and waterfowl. Marine algae supply much of the oxygen required by life on Earth and, at the same time, remove a substantial amount of carbon dioxide from the atmosphere. In addition, the evaporation of the seawater (as part of the water cycle) provides rainwater for the land.

    The ocean can be divided into two general regions: (i) a warm, surface pool—typically 18°C (64°F)—that is approximately 3280 ft thick and (ii) the deep water—typically 3°C (37°F)—that outcrops to the surface at high latitudes and forms the bulk of the ocean volume. Unlike the atmosphere, heating of the ocean surface stabilizes the water and prevents rapid exchange (or mixing) between the surface water and the deeper water. In fact, contact between the surface water and deep water is limited to localized polar regions where the resulting thermohaline circulation (part of the large-scale ocean circulation that is driven by global density gradients that are created by surface heat and freshwater flow) is especially important over long timescales (such as glacial cycles).

    In terms of the chemical effects of the ocean, the ocean influences the atmosphere through the exchanges of gases across the air-sea interface. The transfer of carbon dioxide from the atmosphere to the ocean is controlled by the two competing factors of temperature: warming of surface waters, which releases carbon dioxide to the atmosphere, and biological productivity. Photosynthesis by marine phytoplankton converts dissolved carbon dioxide into organic carbon compounds, leading to a reduction in surface carbon dioxide values and a carbon dioxide flow into the ocean. The amount of carbon dioxide dissolved in seawater is quite large due to its high solubility and its reactivity with water to form carbonic acid and its dissociation products.

    1.2.3 The Geosphere and Terrestrial Biosphere

    The geosphere is that part of the Earth upon which humans live and from which food, minerals, and fuels are extracted. It is divided into layers, which include the solid, iron-rich inner core, molten outer core, and the lithosphere (the upper mantle and the crust). Environmental science (whether it is inorganic or organic science) is most concerned with the lithosphere that extends to depths on the order of 62 miles below the surface of the Earth and comprises two systems: (i) the crust and (ii) the upper mantle. The crust (the outer skin of the Earth) is the layer that is accessible to humans and is extremely thin compared with the diameter of the earth, ranging from 5 to 20 miles thick.

    The biosphere, a subdivision of the geosphere, is the relatively thin zone of air, soil, and water that covers the Earth and is capable of supporting life, and ranges from approximately 6 miles from the deepest floor of the ocean and into the atmosphere. Life in the geosphere is dependent on the energy from the sun energy and on the circulation of heat and essential nutrients.

    Thus, the biosphere (i) is virtually contained by the geosphere and hydrosphere in the very thin layer where these environmental spheres interface with the atmosphere, (ii) strongly influences, and in turn is strongly influenced by, the other parts of the environment, (iii) strongly influences the various parts of the aquasphere by producing the biomass required for life in the water and mediating oxidation-reduction reactions in the water, (iv) is involved with weathering processes that break down rocks in the geosphere and convert the rock matter to soil, and (v) is based upon plant photosynthesis, which fixes solar energy and carbon from atmospheric carbon dioxide in the form of high-energy biomass.

    The anthroposphere is a name given to that part of the environment made

    Enjoying the preview?
    Page 1 of 1